Mechanically Induced Solid-State Generation of Phosphorus Ylides

May 10, 2002 - We describe the nearly quantitative preparation of phosphorus ylides and the Wittig reaction occurring in the solid sate during high-en...
6 downloads 11 Views 39KB Size
Published on Web 05/10/2002

Mechanically Induced Solid-State Generation of Phosphorus Ylides and the Solvent-Free Wittig Reaction Viktor P. Balema,*,† Jerzy W. Wiench,† Marek Pruski,† and Vitalij K. Pecharsky*,†,‡ Ames Laboratory, Iowa State UniVersity, Ames, Iowa 50011-3020, Department of Materials Science and Engineering, Iowa State UniVersity, Ames, Iowa 50011 Received December 31, 2001

Mechanical alloying via high-energy ball-milling is a scalable experimental technique, which is broadly used for the preparation and modification of metals, hydrides, and other inorganic solids.1 In contrast, mechanochemistry is seldom attempted in organic synthesis, and when it is, mechanical processing of organic reactants is followed by additional treatment (usually heating), or it is carried out in the presence of a solvent.2 Consequently, little is known about chemical reactions taking place during mechanical milling of organic solids. Furthermore, only indirect evidence confirming just a few chemical transformations of solid organic compounds as a result of milling has been presented to date.2e,f We show that three major types of phosphorus ylides, that is, stabilized, semistabilized and nonstabilized phosphoranes, can be generated mechanochemically in a solid state. Stabilized ylides were isolated in a pure form, while semistabilized and nonstabilized phosphoranes were trapped by a solid organic carbonyl compound and transformed into ethenes and triphenylphosphine oxide in a solvent-free mechanochemical Wittig-type reaction. Conventionally, phosphorus ylides are produced in a solution,3 and the solid-state preparation of phosphoranes was not reported before. Solid phosphonium salts, 1a-d, and an excess of anhydrous K2CO3 were ball-milled in a hardened-steel vial with steel balls under helium for various periods of time (Scheme 1 and Table 1). Forced air-cooling of the vial was employed to prevent its heating during ball-milling experiments. The obtained powders were analyzed using solid-state 31P nuclear magnetic resonance (NMR) spectroscopy, X-ray powder diffraction (XRD), and differential thermal analysis (DTA) prior to any further treatment. Two stable phosphorus ylides, (benzoylmethylene)triphenylphosphorane (2a) and (carbethoxymethylene)triphenylphosphorane (2b), were prepared in nearly quantitative yields (Scheme 1, Table 1). To prepare ylide 2a, (phenacyl)triphenylphosphonium bromide (1a) and anhydrous K2CO3, taken in a 1:1.2 molar ratio, were ball-milled in a Spex 8000 mill for 1 and 3 h. The solid-state 31P magic angle spinning (MAS) NMR spectrum of the powder sample after ballmilling for 1 h consists of two peaks at δ31P ) 23 and 17 ppm (Figure 1b). The peak at δ31P ) 23 ppm corresponds to the unreacted 1a (Figure 1a) remaining in the reaction mixture. The location of the second peak at δ31P ) 17 ppm is consistent with pure 2a (Figure 1d). After ball-milling for 3 h, the spectrum of the reaction mixture (Figure 1c) contains only one signal at δ31P ) 17 ppm, indicating a complete transformation of 1a into 2a. The sample ball-milled for 1 h was reexamined 6 months after the completion of the mechanical treatment. No continuing changes, that is, no thermal solid-state reactions were detected. Similarly, 2b was * Corresponding authors. E-mail: [email protected], [email protected]. † Ames Laboratory. ‡ Department of Materials Science and Engineering.

6244

9

J. AM. CHEM. SOC. 2002, 124, 6244-6245

Scheme 1. Mechanochemical Generation of Phosphoranes

Table 1. Mechanochemically Prepared Compounds compound

milling time, h

yield, %

E:Z ratioa

referenceb

2a 2b 3a 3b 3c 3d 4 5

3 4 7 8 14 20 4 8

99 96 85d 92d 70d 73d 99 93d

1.6:1e 2:1 3.4:1 3.5:1

c 3b, 8 7 9 10 11 6 7

a E:Z-isomer ratio as determined by the liquid-state (CDCl ) 1H NMR 3 spectroscopy. b References used for identification of the compounds. c Identified by comparison with the commercial phosphorane. d Triphenylphosphine oxide was isolated as a coproduct in the yields matching that of ethene. e Estimated ratio based on the amount of the isolated E-isomer.

prepared mechanochemically from (carbethoxymethyl)triphenylphosphonium bromide (1b) and anhydrous K2CO3, see Table 1. Semistabilized and nonstabilized phosphoranes were generated from (benzyl)triphenylphosphonium bromide (1c) or (methyl)triphenylphosphonium bromide (1d) by mechanical processing with an excess of K2CO3 in the presence of stoichiometric amounts of solid organic carbonyl compounds. The latter served as traps for highly reactive phosphoranes. Thus, we successfully carried out several solvent-free Wittig-type reactions and prepared a series of aromatic ethenes 3a-d (Scheme 1, Table 1) in high yields. Again, the solid-state 31P MAS NMR confirmed the completion of all performed reactions during ball-milling (e.g., Figure 1e-g). Finally, we found that the mechanochemical preparation of stabilized phosphorus ylides and the solvent-free Wittig reaction can be successfully carried out as a “one-pot” process starting with triphenylphosphine, an organic halogenide, and an organic carbonyl compound in the presence of K2CO3 (Scheme 2, Table 1). In this way, we synthesized the ylides 2a and 4, and the bromostilbene 5. It is worth noting that 4 is usually prepared in a solution in a multiple-step process,6 while in our approach the reactants were simply ball-milled together for 4 h. In all cases, the completion of 10.1021/ja017908p CCC: $22.00 © 2002 American Chemical Society

COMMUNICATIONS

Figure 1. Solid-state 31P MAS NMR spectra of: (a) crystalline 1a (Lancaster synthesis); reaction mixtures after ball-milling of 1a with K2CO3 for (b) 1 h and (c) 3 h; (d) crystalline 2a (Alfa Aesar); (e) crystalline 1d (Aldrich); (f) reaction mixture after ball-milling of 1d with K2CO3 in the presence of 2-naphthaldehyde for 10 h (the reaction is essentially complete after 20 h); (g) crystalline triphenylphosphine oxide. Spectra were obtained at 161.9 MHz under single-pulse excitation, 20 kHz MAS and continuous wave 1H decoupling. The more efficient method of crosspolarization from 1H nuclei yielded similar results. Scheme 2. “One-Pot” Mechanochemical Synthesis

mechanochemical transformations was confirmed by the solid-state 31P MAS NMR spectroscopy. Two distinct dissimilarities between the conventional and the mechanochemically induced Wittig reactions of semistabilized and nonstabilized phosphorus ylides are worth noting. First, both (benzylene)- and (methylene)triphenylphosphoranes can be generated mechanochemically using K2CO3, while their preparation in a solution requires much stronger bases.3 As far as we are aware, deprotonation of a (methyl)triphenylphosphonium salt with an alkali metal carbonate was never reported in the literature. Second, while the Wittig reaction between (benzylene)- or (2-naphthylene)triphenylphosphoranes and aromatic carbonyl compounds in a solution produces preferably Z-stilbenes or mixtures with nearly equal content of Z- and E-isomers,3,7 the mechanochemical technique discriminates between Z- and E-substituted products in favor of more thermodynamically stable E-stilbenes. To obtain preliminary insights into the processes resulting in the mechanochemical generation of phosphorus ylides, the powder obtained after ball-milling of the phosphonium salt 1a with K2CO3 for 1 h was studied using DTA and XRD. The major objective was to explore the possibility of local melting of the reactants and, hence, hidden liquid-state reactions during mechanical processing. The data on the local temperature rise in different types of mills during milling of inorganic solids1,4 and our results on mechanical treatment of LiAlH4-based composites5 showed that the materials’ temperature during milling in a Spex 8000 mill do not exceed 110 °C. Since melting points of 1a and anhydrous K2CO3 are 265 and 891 °C, respectively, their melting during mechanical processing is unlikely. On the other hand, the likelihood of the formation of low-melting eutectics from these compounds during ball-milling required further investigation.

Our experiments produced no evidence for melting of the studied material below 180 °C, which is the melting point of the ylide 2a formed during mechanical processing. Furthermore, DTA, XRD, and NMR data (compare Figure 1, c and d) revealed the presence of an amorphous phase in the ball-milled sample. Partial loss of crystallinity was also observed during mechanical treatment of the pure 1a and 2a for 1 h. According to DTA and in situ hightemperature XRD, crystallization of the amorphous components in the investigated powders occurs between 70 and 105 °C, also indicating that the temperature of the reaction mixtures during processing remained below ∼70 °C. The acquired data confirm the solid-state character of the events leading to the mechanochemical generation of phosphorus ylides. They also allow us to explain the latter as a sequence of solid-state processes, which include: (i) breaking the crystal lattice of a phosphonium compound and the formation of an amorphous phase and (ii) a deprotonation of an amorphous phosphonium salt by microcrystalline K2CO3 in a heterogeneous solid-state reaction. In conclusion, mechanically induced solid-state generation of phosphorus ylides and the solvent-free Wittig reaction have been observed for the first time. The completion of the reactions during ball-milling has been directly confirmed by solid-state 31P MAS NMR spectroscopy. High-energy mechanical processing is responsible for both the amorphization of the reactants and mass transfer in a solid state, thus playing a critical role in the discovered chemical transformations. Acknowledgment. This work was supported by the Office of Basic Energy Sciences, Materials Sciences Division (V.P.B. and V.K.P.) and Chemical Sciences Division (J.W. and M.P.) of the U.S. DOE. The authors thank to Mr. Kevin Dennis and Mr. Keita Hosokawa for help with DTA measurements, and to Dr. Karl A. Gschneidner, Jr., for discussions of the results. Supporting Information Available: Experimental details, DTA and XRD data (PDF). This material is available free of charge via the Internet at http://pubs.acs.org. References (1) (a) Suryanarayana, C. Prog. Mater. Sci. 2001, 46, 1. (b) Volkov V. V.; Myakishev K. G.; Inorg. Chim. Acta 1999, 289, 51. (2) (a) Toda, F. Acc. Chem. Res. 1995, 28, 8, 480. (b) Nu¨chter, M.; Ondruschka, B.; Trotzki, R. J. Prakt. Chem. 2000, 342, 720. (c) Makhaev, V. D.; Borisov, A. P.; Petrova, L. A. J. Organomet. Chem. 1999, 590, 222. (d) Tanaka, K.; Toda, F. Chem. ReV. 2000, 100, 1025. (e) Murata, Y.; Kato N.; Komatsu, K. Chem. Commun. 2000, 1583. (f) Komatsu, K.; Murata, Y.; Wang G.-W.; Tanaka, T.; Kato N.; Fujieara K. Fullerene Sci. Technol. 1999, 7, 609. (3) (a) Johnson, W. A. Ylides and Imines of Phosphorus; John Wiley & Sons: New York, 1993. (b) Kolodiazhny, O. I. Phosphorus Ylides; WileyVCH: Weinheim, 1999. (c) Vedejs, E.; Peterson, M. J. Top. Stereochem. 1994, 21, 1. (4) Koch, C. C. Int. J. Mechanochem. Mech. Alloying 1994, 1, 56. (5) (a) Balema, V. P.; Dennis, K. W.; Pecharsky, V. K. J. Alloys Compds. 2000, 313, 69. (b) Balema, V. P.; Wiench, J. W.; Pruski, M.; Pecharsky, V. K. J. Alloys Compds. 2001, 329, 108. (6) (a) Aitken, A. R.; Cadogan, J. I. G.; Gosney, I. Phosphorus, Sulfur Silicon Relat. Elem. 1995, 101, 281. (b) Brittain, J. M.; Jones, R. A. Tetrahedron 1979, 35, 1139. (7) Brooks, M. A.; Scott, L. T. J. Am. Chem. Soc. 1999, 121, 5444. (8) (a) Climent, M. S.; Marinas, J. M.; Mouloungui, Z.; Le Bigot, Y.; Delmas, M.; Gaset, A.; Sinisterra, J. V. J. Org. Chem. 1989, 54, 3695. (b) Weleski, E. T.; Silver, J. L.; Jansson, M. D.; Burmeister, J. L. J. Organomet. Chem. 1975, 102, 365. (9) Kumari, N.; Kendurkar, P. S.; Tewari, R. S. J. Organomet. Chem. 1975, 96, 237. (10) Beller, M.; Riermeier, T. H. Eur. J. Inorg. Chem. 1998, 29. (11) (a) Katritzky, A. R.; Hitchings, G. J.; King, R. W.; Zhu, D. W. Magn. Reson. Chem. 1991, 29, 2. (b) Ando, W.; Sekiguchi, A. J. Organomet. Chem. 1977, 133, 219.

JA017908P

J. AM. CHEM. SOC.

9

VOL. 124, NO. 22, 2002 6245